首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A three-dimensional potential energy surface for the 2A′ ground state of the system (Ne? H2)+ (2Σ+ in collinear geometry) has been calculated at SCF and CEPA levels. This surface describes the abstraction reaction which is endoergic by 0.57 eV (ΔH00) and has been studied recently by different experimental groups at low collision energies. Our CEPA calculations yield an endoergicity of 0.55 eV (ΔH00). The 2A′ surface has a minimum at collinear geometry with RNe—H = 2.29 a0 and RH? H = 2.08 a0 and a well depth of 0.49 eV relative to Ne + H+2. The effects of electron correlation on the shape of the surface and on the well depth are discussed. An analytic fit of the collinear part of the surface has been constructed based on Simon's proposal of using polynomials in the coordinates (R? Re)/R instead of (R? Re). The fitted potential is used for quantum mechanical scattering calculations with the finite element method (FEM ). Preliminary results for reaction probabilities for H+2 in different vibrationally excited states are given and compared to the experimental results.  相似文献   

2.
The kinetics of the redox reaction between mandelic acid (MA) and ceric sulfate have been studied in aqueous sulfuric acid solutions and in H2SO4? MClO4 (M+ = H+, Li+, Na+) and H2SO4? MHSO4 (M+ = Li+, Na+, K+) mixtures under various experimental conditions of total electrolyte concentration (that is, ionic strength) and temperature. The oxidation reaction has been found to occur via two paths according to the following rate law: rate = k[MA] [Ce(IV)], where k = k1 + k2/(1 + a)2[HSO4?]2 = k1 + k2/(1 + 1/a)2[SO42?]2, a being a constant. The cations considered exhibit negative specific effects upon the overall oxidation rate following the order H+ ? Li+ < Na+ < K+. The observed negative cation effects on the rate constant k1 are in the order Na+ < Li+ < H+, whereas the order is in reverse for k2, namely, H+ ? Li+ < Na+. Lithium and hydrogen ions exhibit similar medium effects only when relatively small amounts of electrolytes are replaced. The type of the cation used does not affect significantly the activation parameters.  相似文献   

3.
The low-energy fragmentation characteristics of the [H3,C,N,O2] isomers [H3CNO2] (a), [H2C?N(O)OH] (b), [H3CONO] (c), [HC(O)NHOH] (d) and [HC(OH)?NOH] (e) were studied in detail by metastable ion mass spectrometry. In agreement with most earlier observations, appearance energy measurements established the potential energy surface of the isomers a, b and c, showing the intricate interrelations between them. It was concluded that a isomerizes into b prior to fragmentation by loss of ·OH and H2O and into c before loss of ·H and H3CO· moreover, the reverse reactions do not take place on the metastable time-frame. The dominant metastable process for isomers d and e (obtained via HCN loss from glyoxime) was generation of [H2NOH]. For isomer e this process was proposed to involved a rate-determining isomerization into d. It was concluded that isomers d and e do not intercommunicate with ions a, b and c prior to fragmentation. Neutralization-reionization mass spectrometry indicated that the enol form of formohydroxamic acid as well as the keto counterpart are stable in the gas phase.  相似文献   

4.
The formation of 1 : 2 titanium(III) complex with chromotropic acid (4, 5-dihydroxy-2, 7-naphthalene-disulfonic acid) was observed by spectrophotometric measurements at various ionic strengths. An expression, [Ti(III)]/D=1/Δ? + αH2+/KΔ?[H2R2?]2, was derived for the determination of the formation constant, K=7.2×102 liter2 mol?2 for the Ti(III).(HR)2 ion in the pH range of 1.3–1.8 at constant ionic strength, I=0.2 M, at 25°C. The thermodynamic data for the reaction, Ti(III)+2H3R2?=Ti(III) (HR)2+2H+, were calculated to be ΔG° = ?16 kJ mol?1 ΔH° = 18 kJ mol?1, ΔS° = 110 JK?1 mol?1, at 25°C.  相似文献   

5.
We report that 2,6‐lutidine?trichloroborane (Lut?BCl3) reacts with H2 in toluene, bromobenzene, dichloromethane, and Lut solvents producing the neutral hydride, Lut?BHCl2. The mechanism was modeled with density functional theory, and energies of stationary states were calculated at the G3(MP2)B3 level of theory. Lut?BCl3 was calculated to react with H2 and form the ion pair, [LutH+][HBCl3?], with a barrier of ΔH=24.7 kcal mol?1G=29.8 kcal mol?1). Metathesis with a second molecule of Lut?BCl3 produced Lut?BHCl2 and [LutH+][BCl4?]. The overall reaction is exothermic by 6.0 kcal mol?1rG°=?1.1). Alternate pathways were explored involving the borenium cation (LutBCl2+) and the four‐membered boracycle [(CH2{NC5H3Me})BCl2]. Barriers for addition of H2 across the Lut/LutBCl2+ pair and the boracycle B?C bond are substantially higher (ΔG=42.1 and 49.4 kcal mol?1, respectively), such that these pathways are excluded. The barrier for addition of H2 to the boracycle B?N bond is comparable (ΔH=28.5 and ΔG=32 kcal mol?1). Conversion of the intermediate 2‐(BHCl2CH2)‐6‐Me(C5H3NH) to Lut?BHCl2 may occur by intermolecular steps involving proton/hydride transfers to Lut/BCl3. Intramolecular protodeboronation, which could form Lut?BHCl2 directly, is prohibited by a high barrier (ΔH=52, ΔG=51 kcal mol?1).  相似文献   

6.
We report sufficient theoretical evidence of the energy stability of the e+?H22? molecule, formed by two H? anions and one positron. Analysis of the electronic and positronic densities of the latter compound undoubtedly points out the formation of a positronic covalent bond between the otherwise repelling hydride anions. The lower limit for the bonding energy of the e+?H22? molecule is 74 kJ mol?1 (0.77 eV), accounting for the zero‐point vibrational correction. The formation of a non electronic covalent bond is fundamentally distinct from positron attachment to stable molecules, as the latter process is characterized by a positron affinity, analogous to the electron affinity.  相似文献   

7.
Dissociation equilibria of lithiumthiocyanate (LiSCN) in N,N-dimethylformamide (DMF) solutions of poly (propylene oxide) (PPO) were investigated by using infrared spectroscopy. The stretching bands due to the thiocyanate ions SCN?1 and the LiSCN ion pair were found at 2058 and 2072 cm?1, respectively. At high LiSCN concentration C of ca. 20 wt %, another weak band due to the dimer (LiSCN)2 was observed. From the ratio of the areas of the absorption bands, the dissociation constant K1 for the equilibrium LiSCN ? Li+ + SCN?1 and that K2 for (LiSCN)2 ? 2LiSCN have been determined. With increasing DMF content, K1 increases from 1 × 10?4 for bulk PPO to 4.8 × 10?1 for pure DMF at 299 K. Log K1 is not linear against inverse of the dielectric constant ? of the medium and decreases with increasing temperature. The enthalpy (ΔH) and entropy (ΔS) changes for the dissociation of LiSCN are both negative. ©1995 John Wiley & Sons, Inc.  相似文献   

8.
The ion [C13H9]+ (m/e 165) is produced from the molecular ion of 3,5-diphenylisoxazole by the process [M ? CO ? H2CN·] and [M ? CO ? HCN ? H·] and from that of 3,5-diphenyl-pyrazole by the eliminations [M ? N2H· ? C2H2]. These processes have been studied by 2H and 13C labelling. A correlation between photochemical, thermal and electron-impact decompostions is noted for 3,5-diphenylisoxazole.  相似文献   

9.
The xenon–difluoronitrenium ion F2N? Xe+, a novel xenon–nitrogen species, was obtained in the gas phase by the nucleophilic displacement of HF from protonated NF3 by Xe. According to Møller–Plesset (MP2) and CCSD(T) theoretical calculations, the enthalpy and Gibbs energy changes (ΔH and ΔG) of this process are predicted to be ?3 kcal mol?1. The conceivable alternative formation of the inserted isomers FN? XeF+ is instead endothermic by approximately 40–60 kcal mol?1 and is not attainable under the employed ion‐trap mass spectrometric conditions. F2N? Xe+ is theoretically characterized as a weak electrostatic complex between NF2+ and Xe, with a Xe? N bond length of 2.4–2.5 Å, and a dissociation enthalpy and free energy into its constituting fragments of 15 and 8 kcal mol?1, respectively. F2N? Xe+ is more fragile than the xenon–nitrenium ions (FO2S)2NXe+, F5SN(H)Xe+, and F5TeN(H)Xe+ observed in the condensed phase, but it is still stable enough to be observed in the gas phase. Other otherwise elusive xenon–nitrogen species could be obtained under these experimental conditions.  相似文献   

10.
Dissociative electron ionization (70eV) of selenophene (C4H4Se) generates m/z 106 ions of composition [H2, C2, 80Se]+? and m/z 105 ions of [H, C2, 80Se]+. From tandem mass spectrometric experiments, Density Functional Theory (DFT) and ab initio calculations, it is concluded that these ions have the structure of selenoketene H2C?C?Se+? (1a+? )and selenoketyl HC?C?Se+ (2a+) ions respectively. The calculations predict that selenoketene ion 1a+? is separated by high energy barriers from its isomers selenirene (H e)+? 1b+?, ethyne selenol (HCCSeH)+? 1c+?, (CCHSeH)+? 1d+? and (CCSeH2)+? 1e+?. The selenoketyl ion 2a+ is separated by high barriers from its isomers (CCHSe)+ 2b+, and (CCSeH)+ 2c+. Neutralization‐reionization mass spectra (NRMS) of these structurally characterized ions confirmed that the corresponding neutral analogues, selenoketene H2CCSe 1a and selenoketyl radical HCCSe 2a? are stable in the rarefied gas phase. The relative, dissociation, and isomerization energies for selenoketene and selenoketyl ions and neutrals studied at B3LYP/6–31G(d,p) and G2/G2(MP2) levels are used to support and interpret the experimental results. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

11.
The genesis of m/e 147 [C5H15OSi2]+ from the title compounds is dependent upon the geometry of the transition state of the reaction [M ? CH3]+m/e 147.  相似文献   

12.
The kinetics of stripping of Ni2+ from a Ni‐BTMPPA complex, dissolved in a kerosene solution of BTMPPA (H2A2, Cyanex 272), by acidic sulfate‐acetato solution, was studied using the single (falling) drop technique and flux (F) method of data treatment. The empirical flux equation at 303 K is Fb (kmol/m2s) = 10?4.35 [Ni2+] (1+10?3.42 [H+]?1)?1 ([H2A2](o)0.5+2.50 [H2A2](o))?1 (1+6[SO42?]) (1+3.20 [Ac?]). Activation energy (Ea), entropy change in activation (ΔS±), and enthalpy change in activation (ΔH±) were measured under different experimental conditions. Based on the empirical flux equation, Ea and ΔS±, the mechanism of Ni2+ stripping is provided. In a low [H+] region, the stripping reaction steps appear as [NiA+] → Ni2+ + A? and [Ni(HA2)2](int) → [NiHA2]+(int) + HA2(int)? in lower and higher concentration regions of free BTMPPA, respectively, provided [SO42?] and [Ac?] are kept low. However, at higher [H+] concentrations, the stripping is under diffusion control. With increasing [SO42?] and [Ac?], the enhancement of the rate is attributed to the attack of the Ni(II) complex by SO42? or HSO4? and Ac? to form NiSO4 or NiHSO4+ and NiAc+ complexes. Negative ΔS± values indicate that the rate‐determining stripping reaction steps occur via an substitution nucleophilic, bimolecular (SN2) mechanism.  相似文献   

13.
The characteristics of the in-beam electron impact mass spectra of six isomers of undecanol as well as several 1-alkanols have been examined. In addition to the characteristic ions observed in the conventional electron impact spectra, the [2M+1]+, [2M+1-H2O]+, [2M+1-2H2O]+, [2M-R or R′]+, [2M-H2O? R or R′]+, [2M? 2H2O? R or R′]+ and [M+1? H2O]+ peaks are common in the in-beam electron impact mass spectra of the undecanol isomers of structure RCH(OH)R′. Deuterium labelling experiments have shown that the extra proton in the protonated dimer ions, [2M+1]+, originates from the hydroxy group. The processes which produce the important peaks in the high m/e regions are discussed.  相似文献   

14.
The stabilization energies (ΔEform) calculated for the formation of the Li+ complexes with mono‐, di‐ tri‐ and tetra‐glyme (G1, G2, G3 and G4) at the MP2/6‐311G** level were ?61.0, ?79.5, ?95.6 and ?107.7 kcal mol?1, respectively. The electrostatic and induction interactions are the major sources of the attraction in the complexes. Although the ΔEform increases by the increase of the number of the O???Li contact, the ΔEform per oxygen atom decreases. The negative charge on the oxygen atom that has contact with the Li+ weakens the attractive electrostatic and induction interactions of other oxygen atoms with the Li+. The binding energies calculated for the [Li(glyme)]+ complexes with TFSA? anion (glyme=G1, G2, G3, and G4) were ?106.5, ?93.7, ?82.8, and ?70.0 kcal mol?1, respectively. The binding energies for the complexes are significantly smaller than that for the Li+ with the TFSA? anion. The binding energy decreases by the increase of the glyme chain length. The weak attraction between the [Li(glyme)]+ complex (glyme=G3 and G4) and TFSA? anion is one of the causes of the fast diffusion of the [Li(glyme)]+ complex in the mixture of the glyme and the Li salt in spite of the large size of the [Li(glyme)]+ complex. The HOMO energy level of glyme in the [Li(glyme)]+ complex is significantly lower than that of isolated glyme, which shows that the interaction of the Li+ with the oxygen atoms of glyme increases the oxidative stability of the glyme.  相似文献   

15.
By ion cyclotron resonance it is found that various substituted cy clopropanes after ionization react with ammonia to give products which allow identification of the degree and kind ofsubstitution on the cyclopropl ring. For example, cyclopropyle reacts to give [CH2NH2]+ (m/e 30), methylcyclopropane gives [CH2NH2]+ (m/e 30) and the ethyl substituted [CH(C2H3)NH2]+ (m/e 44) and ethylopropane gives [CH2NH2]+ (m/e 30) and the ethyl substituted [CH(C2H5)NH2]+ (m/e 58). It is suggested that reactions of stable molecular ions with reagent neutrals may be a source of highly specific structural information for organic compounds.  相似文献   

16.
Low energy collision induced dissociation (CID) spectra were measured by a triple stage quadrupole mass spectrometer for the [MH]+ ions of diethyl and dimethyl esters of maleic, fumaric, citraconic and mesaconic acids. A very high degree of stereospecificity was observed for the geometrically isomeric diethyl esters. The cis esters give rise to very abundant [MH? EtOH]+ and [MH? EtOH? C2H4]+ ions, while the trans isomers exhibit very abundant [MH? C2H4]+ and [MH? 2 C2H4]+ ions. The highly stereospecific processes indicate that the double bond configuration is retained in the protonated species under the conditions of the experiment.  相似文献   

17.
Unstable 2-hydroxpropene was prepared by retro-Diels-Alder decomposition of 5-exo-methyl-5-norbornenol at 800°C/2 × 10?6 Torr. The ionization energy of 2-hydroxypropene was measured as 8.67±0.05 eV. Formation of [C2H3O]+ and [CH3]+ ions originating from different parts of the parent ion was examined by means of 13C and deuterium labelling. Threshold-energy [H2C?C(OH)? CH3] ions decompose to CH3CO++CH3˙ with appearance energy AE(CH3CO+) = 11.03 ± 0.03 eV. Higher energy ions also form CH2?C?OH+ + CH3 with appearance energy AE(CH2?C?OH+) = 12.2–12.3 eV. The fragmentation competes with hydrogen migration between C(1) and C(3) in the parent ion. [C2H3O]+ ions containing the original methyl group and [CH3]+ ions incorporating the former methylene and the hydroxyl hydrogen atom are formed preferentially, compared with their corresponding counterparts. This behaviour is due to rate-determining isomerization [H2C?C(OH)? CH3] →[CH3COCH3], followed by asymmetrical fragmentation of the latter ions. Effects of internal energy and isotope substitution are discussed.  相似文献   

18.
Third-order Møller–Plesset perturbation theory (MP 3) with a 6-31G** basis set was applied to study the relative stabilities of H+(X)2 conformations (X ? CO and N2) and their clustering energies. The effect of both basis set extensions and electron correlation is not negligible on the relative stabilities of the H+(CO)2 clusters. The most stable conformation of H+(CO)2 is found to be a Cv structure in which a carbon atom of CO bonds to the proton of H+(CO), whereas that of H+(N2)2 is a symmetry Dh structure. The second lowest energy conformations of H+(CO)2 and H+(N2)2 lie within 2 kcal/mol above the energies of the most stable structures. Clustering energies computed using MP 3 method with the 6-31G** basis set are in good agreement with the experimental findings of Hiraoka, Saluja, and Kebarle. The low-lying singlet conformations of H+(X)3 (X ? CO and N2) have been studied by the use of the Hartree–Fock MO method with the 6-31G** basis set and second-order Møller–Plesset perturbation theory with a 4-31G basis set. The most stable structure is a T-shaped structure in which a carbon atom of CO (or a nitrogen atom of N2) attacks the proton of the most stable conformation of H+(X)2 clusters.  相似文献   

19.
Protonated silanoic acid, HSi(OH)2+, 1a +, is cleanly generated by the dissociative electron ionization of triethoxysilane, HSi(OC2H5)3, and tetraethoxysilane, Si(OC2H5)4. This follows from tandem mass spectrometric experiments and CBS-QB3 model chemistry calculations. The calculations predict that 1a +Hf(298 K) = 205 kJ mol−1) is separated by high barriers from its isomers HOSiOH2+, 1b + and HSi(O)OH2+, 1c +. Low-energy (metastable) ions 1a + dissociate by loss of H2O via the pathway 1a + → 1b + → SiOH+ + H2O. Analysis of the metastable peak for this process confirms that the isomerization step 1a + → 1b + is rate determining. The calculations further predict that the incipient ions 1b + communicate via a low barrier with the proton-bound dimer SiO···H···OH2+, 1d +. This dimer ion is much lower in energy than its counterpart OSi···H···OH2+, 1e +, which is calculated to be only marginally stable. A comparison of the potential energy diagram for the silicon-containing ions 1a +– 1e + with that of their carbon analogues reveals that the dissociation chemistries of HSi(OH)2+ and HC(OH)2+ are only superficially similar. Neutralization–reionization experiments confirm the theoretical prediction that the HSi(OH)2 radical (ΔHf(298 K) = −455 kJ mol−1) is a stable species in the rarefied gas phase. However, owing to a mismatch of Franck–Condon factors a large fraction of the neutralized ions dissociates by loss of H yielding Si(OH)2. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

20.
Experimental relative abundances of D2O+, OD+ and D+ ions are given as functions of the initial energy of D2O+ ions in the B?2B2 state. The D+ ions are more abundant than expected on the basis of a recent theoretical model. Less kinetic energy is released in D+ formation than in OD+ formation, for the same excess energy available. It is suggested that OD+ formation should be modelled as a statistical reaction, whereas D+ formation is more specific. No detailed model can be suggested on present evidence, but it is pointed out that H2O+ (B?2B2) can be predissociated to OH+ by the X? state to H+ by the à state, as well as by repulsive states.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号